首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Electron transport in immobilized liquid membranes was studied in the reagent concentration-dependent regime. The velocity dependence of cells utilizing Vitamin K3 (VK3) and 2-tert-butylanthraquinone (TBAQ) as carriers was determined. The velocity for the VK3 cell was proportional to the product [MV+]0.5[VK3]0.9 [Fe(phen33+]0.5 exp 40 kJ/RT; that of the TBAQ cell was proportional to the product [MV+]1.3 [TBAQ]1.5 [Fe(phen)33+]0.9 exp 10 kJ/RT. The velocity of electron transport was kinetically controlled at the oxidant interface as verified by independent rate measurements. The rate constant for the reaction of MV+ with TBAQ was 8.4 x 108 M−1-sec−1 (reductant interface); for the reaction of H2TBAQ with Fe(phen)33+, it was 34 M−1-sec−1(oxidant interface). The velocity dependence reduced to the following in the concentration independent regime: for the TBAQ cell, v ∞ exp 10 kJ/RT ([MV+] >0.6 mM, [TBAQ] >0.2 M, [Fe(phen)33+] > 5 mM, [Ru(bpy)32+] > 0.2 mM); for the VK3 cell, v ∞ [VK3] exp 40 kJ/RT ([MV+] > 0.4 mM, [Fe(phen)33+] > 5 mM, [Ru(bpy)33+] > 0.2 mM). The mechanism of electron transport in the TBAQ cell is best interpreted to involve formation of semiquinone and hydroquinone in the membrane which then react with Fe(phen)33+ in the rate-limiting electron transfer step.  相似文献   

2.
We have identified compounds obtained from the SARA fractions of bitumen by using atmospheric pressure photoionization mass spectrometry and low‐energy collision tandem mass spectrometric analyses with a QqToF‐MS/MS hybrid instrument. The identified compounds were isolated from the maltene saturated oil and the aromatic fractions of the SARA components of a bitumen. The QqToF instrument had sufficient mass resolution to provide accurate molecular weight information and to enhance the tandem mass spectrometry results. The APPI‐QqToF‐MS analysis of the separated compounds showed a series of protonated molecules [M + H]+ and molecular ions [M]+? of the same mass but having different chemical structures, in the maltene saturated oil and the aromatic SARA fractions. These isobaric ions were a molecular ion [M2]+? at m/z 418.2787 and a protonated molecule [M5 + H]+ at m/z 287.1625 in the saturated oil fraction, and molecular ions [M6]+? at m/z 418.1584 and [M7]+? at m/z 287.1285 in the aromatic fraction. The identification of this series of chemical compounds was achieved by performing CID‐MS/MS analyses of the molecular ions [M]+? ([M1]+? at m/z 446. 2980, [M2]+? at m/z 418.2787, [M3]+? at m/z 360.3350 and [M4]+? at m/z 346.2095) in the saturated oil fraction and of the [M5 + H]+ ion at m/z 287.1625 also in the saturated oil fraction. The observed CID‐MS/MS fragmentation differences were explained by proposed different breakdown processes of the precursor ions. The presented tandem mass spectrometric study shows the capability of MS/MS experiments to differentiate between different classes of chemical compounds of the SARA components of bitumen and to explain the reasons for the observed mass spectrometric differences. However, greater mass resolution than that provided by the QqToF‐MS/MS instrument would be required for the analysis of the asphaltene fraction of bitumen. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

3.
Curcumin, a polyphenolic compound, has shown a wide range of pharmacological activities and has been widely used as a food additive. However, the clinical use of curcumin is limited to some extent because of its poor water solubility and low bioavailability. To overcome these problems, many approaches have been attempted and structural modification of curcumin by microbial transformation has been proven to be an alternative. In this study, we isolated a novel yeast strain Pichia kudriavzevii ZJPH0802 from a soil sample, which is capable of converting curcumin to its derivatives. The transformed products by this strain were evaluated by HPLC, (+) electrospray ionization (ESI)-MSn, and 1H nuclear magnetic resonance methods. Compared with controls, two new peaks of the transformed broth appeared at retention times of 26 min (I) and 62 min (II) by HPLC analysis. The two transformed products were then further identified by (+) ESI-MSn. The spectrum showed that compound I had an accurate [M+H+NH3]+ ion at m/z 392, [M+H]+ ion at m/z 375, [M+H–H2O]+ ion at m/z 357, and (+) ESI-MS3 spectrum showed that ion at m/z 357 could further form fragment ions at m/z 339, 177, and 163; compound II had an accurate [M+H]+ ion at m/z 373, [M+H–H2O]+ ion at m/z 355, and (+) ESI-MS3 spectrum showed that ion at m/z 355 could further form fragment ions at m/z 219, 179, 177, 163, and 137. These two transformed products thereby were confirmed as hexahydrocurcumin (I) and tetrahydrocurcumin (II).  相似文献   

4.
The sensitivity of detection of uric acid (H2U) in positive ion mode electrospray ionization mass spectrometry (ESI MS) was enhanced by uric acid oxidation during electrospray ionization. With a carrier solution of pH 6.3>pKa1=5.4 of H2U, protonated unoxidized uric acid [H2U+H]+ (m/z 169) was detected together with the protonated uric acid dimer [2H2U+H]+ (m/z 337). The dimer likely forms by 1e? oxidation of urate (HU?) followed by rapid radical dimerization. A covalent structure of the dimer was verified by H/D exchange experiments. Efficiency of 2e?, 2H+ oxidation of uric acid is low during ESI in pH 6.3 carrier solution and improves when a low on‐line electrochemical cell voltage is floated on the high voltage of the ES in on‐line electrochemistry ESI MS (EC/ESI MS). The intensity of the uric acid dimer decreases with an increase in the low applied voltage. In a carrier solution with 0.1 M KOH, pH 12.7>pKa2=9.8 of H2U, allantoin (Allnt) (MW 158.04), the final 2e?, 2H+ oxidation product of uric acid, was detected as a potassium complex [K(Allnt)+K]+ (m/z 235) and the [2H2U+H]+ dimer was not detected. In direct ESI MS analysis of 1000‐fold diluted urine [NaHU+H]+ (pKsp NaHU=4.6) was detected in 40/60 (vol%) water/methanol, 1 mM NH4Ac, pH ca. 6.3 carrier solution. A new configuration of the ESI MS instrument with a cone‐shaped capillary inlet significantly enhanced sensitivity in ESI and EC/ESI MS measurements of uric acid.  相似文献   

5.
The mass spectra of some α-substituted phenyl-α,α′-dimethoxyl ketones (compounds 1) and their 2,4-dinitrophenylhydrazones (compounds 2) and semicarbazones (compounds 3) have been studied. The characteristic fragments at m/z (M ? 73) from compounds 1, m/z (M ? 253) from compounds 2 and m/z (M ? 130) from compounds 3 are abundant and proposed to be [ArCROCH3]+. Fragmentations yielding [M+ ? 49] from compounds 2 are abnormal and probably involve the methoxyl and nitro groups. The intense peak at m/z 130 due to [CH3OCH2CNNHCONH2]+ from compounds 3 corresponds to α-cleavage of the molecular ion. Some other fragments from these new compounds are interpreted in this paper.  相似文献   

6.
The IR emission spectra of the molten systems NaCl-CsCl-Cs2CO3-MCl n (M = Li, Be) have been obtained. Spectral data shows that, at a definite limiting molar ratio n lim = [CO3]/[Mn+] characteristic for each Mn+, carbonate-chloride melts based on the NaCl-CsCl eutectic contain carbonate complexes [Li(CO3)3]5? and [Be(CO3)3]4? and, if n < n lim, carbonato chloro complexes [M(CO3) m Cl4?m ], where m = 1–2.  相似文献   

7.
High signal intensities of glutathione (GSH), [GSH+H]+ (m/z 308), cysteine (CySH), [CySH+H]+ (m/z 122), and homocysteine (hCySH), [hCySH+H]+ (m/z 136), are observed in ESI MS with on‐line electrochemistry (EC). Dimers formed by H‐bonding, which are not electrochemical products, are detected as [2GSH+H]+ (m/z 615), [2CySH+H]+ (m/z 243) and [2hCySH+H]+ (m/z 271) together with disulfide dimers GSSG, CySSCy and hCySSCyh, [GSSG+H]+ (m/z 613), [CySSCy+H]+ (m/z 241) and [hCySSCyh+H]+ (m/z 269). When dopamine is present a thiol/dopamine quinone (DAQ) adduct is observed. Formation of this adduct is proposed to occur by an electrochemical mechanism during ESI. Catalysis of thiol oxidation and analysis of thiol mixtures is addressed.  相似文献   

8.
In studying the metabolic pathways underlying the mechanism of carcinogenesis of the heterocyclic amine of 2‐amino‐3‐methylimidazo[4,5‐f]quinoline (IQ), we recently found a new metabolite which gave an [M + H]+ ion of m/z 217 when subjected to electrospray ionization (ESI) in positive‐ion mode. Following ip injection of this metabolite of m/z 217 (designated as m/z 217) to beta‐naphthoflavone‐treated mice, 57% of the total radioactivity was recovered in a 24‐h mouse urine sample. HPLC separation followed by MS analysis indicates that the urine sample contained m/z 217 (36 ± 3% of total recovered radioactivity) and two other peaks that gave rise to the [M + H]+ ions of m/z 393 (31 ± 4%, designated as m/z 393) and m/z 233 (14 ± 1%, designated as m/z 233). Beta‐glucuronidase treatment of m/z 393 resulted in a radioactive peak corresponding to m/z 217. ESI in combination with various mass spectrometry techniques, including multiple‐stage mass spectrometry, exact mass measurements and H/D exchange followed by tandem mass spectrometry, was used for structural characterization. The urinary metabolites of m/z 217, 393 and 233 were identified as 1,2‐dihydro‐2‐amino‐5‐hydroxy‐3‐methylimidazo[4,5‐f]quinoline, 1,2‐dihydro‐2‐amino‐5‐O‐glucuronide‐3‐methylimidazo[4,5‐f]quinoline and 1,2‐dihydro‐2‐amino‐5,7‐dihydroxy‐3‐methylimidazo[4,5‐f]quinoline, respectively. Our results demonstrated that m/z 217 is biotransformed in vivo to m/z 393 by O‐glucuronidation and to m/z 233 by oxidation. The observation of these more polar metabolites relative to IQ suggests that they may arise from a previously undescribed detoxicification pathway. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

9.
This paper describes the characterization of low molecular weight components of four materials using electrospray ionization Fourier transform mass spectrometry (ESI-FTMS). The materials in the current study are [(ViMe2SiO1/2)x(PhSiO3/2)y(SiO4/2)z] (MTQ), [(ViMe2SiO1/2)x(SiO4/2)y] (MQ), and [(SiO4/2)x(HO1/2)y(tBuO1/2)z] (Q) silsesquioxanes. Accurate mass measurements coupled with knowledge of resin chemistry afforded siloxane composition determination that was used to propose specific structures for the oligomers. Branched or linear (TnQmMn+2m+2), and monocyclic (TnQmMn+2m) structures are predominant structures for the low molecular weight species in MTQ. For MQ and Q, more condensed structures, such as partially opened cage structures (QmM2m?6 and QmM2m?8), were identified. The differences between MQ, Q, and MTQ are likely attributed to differences in intrinsic structure and reactivity of T and Q building blocks. The structural information obtained for these oligomeric species will ultimately provide a better understanding of new resin materials and their associated physical properties.  相似文献   

10.
Synergetic extraction of [RuNO(NO2)4OH]2? by calix[4]arene phosphine oxides (L) in the form of Ru/M heterometallic complexes was studied in the presence of nonprecious metals (M2+). The main extraction laws were recognized for [M(NO3)2L n ] and [RuNO(NO2)4OH])ML m ], where M2+ = Zn2+, Cu2+, Co2+, or Ni2+ and n, m = 1 or 2; extraction constants were determined for these metals. The variation row of the extraction constants with varying metal (Zn2+ > Cu2+ > Co2+ > Ni2+) coincides with the Irving-Williams row. Two or three PO groups of extractant L and the OH and NO2 groups of the ruthenium anion are coordinated to the M2+ atom in Ru/M complexes. The conditions for generation of the Ru/Zn complex and its complete extraction were optimized as applied to the extraction of fission ruthenium from nitrated nitric acid and imitation solutions.  相似文献   

11.
The mass spectra of diethyl phenyl phosphates show substituent effects with electron-donating groups favouring the molecular ion M+˙, and the [M? C2H4]+˙, [M – 2C2H4]+˙ and [XPhOH]+˙ ions. The [PO3C2H6]+ (m/z 109) and [PO3H2]+ (m/z 81) ions are favoured by electron-withdrawing groups. Results suggest that the formation of the [XPhC2H3]+˙ ion involves rearrangement of C2H3 to the position ortho to the phosphate group. Ortho effects are also observed.  相似文献   

12.
In this study, a new LC‐ESI‐MS/MS‐based method was validated for the quantitation of hemslecin A in rhesus monkey plasma using otophylloside A as internal standard (IS). Hemslecin A and the IS were extracted from rhesus monkey plasma using liquid–liquid extraction as the sample clean‐up procedure, and were subjected to chromatography on a Phenomenex Luna CN column (150 × 2.0 mm, 3.0 µm) with the mobile phase consisting of methanol and 0.02 mol/mL ammonium acetate (55:45, v/v) at a flow rate of 0.2 mL/min. Detection was performed on an Agilent G6410B tandem mass spectrometer by positive ion electrospray ionization in multiple reaction monitoring mode, monitoring the transitions m/z 580.5 [M + NH4]+ → 503.4 and m/z 518.2 [M + NH4]+ → 345.0 for hemslecin A and IS, respectively. The assay was linear over the concentration range of 0.5–200 ng/mL and was successfully applied to a pharmacokinetic study in rhesus monkeys. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

13.
Dopamine [DA]+ (m/z 154), DA dimer [2DA‐H]+ (m/z 307) and DA quinone [DAQ]+ (m/z 152) are detected in positive ion mode electrospray ionization mass spectrometry (ESI MS) of dopamine in 50/1/49 (vol%) water/acetic acid/methanol. H/D exchange experiments support a covalent structure of DA dimer. Thus, ESI of DA may involve 1e?, 1H+ oxidation processes followed by rapid radical dimerization. The DA quinone signal is low in ESI MS, which indicates a low efficiency of the 2e?, 2H+ oxidation reaction. On‐line electrochemistry ESI MS (EC/ESI MS) with low electrochemical cell voltage floated on high ES voltage increases electrospray current and improves sensitivity for DA. The DA quinone signal increases and DA dimer signal decreases. A new configuration of the ESI MS instrument with a cone‐shaped capillary inlet significantly enhanced sensitivity of ESI and EC/ESI MS measurements. A DA quinone‐cysteine adduct [DAQ+Cys]+ was detected in solutions of DA with cysteine (Cys). ESI MS and EC/ESI MS indicate formation of the DA quinone‐cysteine adduct by 1e? pathway. Oxidation pathways in ESI MS are relevant to biological reactivity of DA and Cys.  相似文献   

14.
The rate of the reaction
has been investigated at 40–65°C with [HClO4] varying from 0.04 to 0.6 M (μ = 0.6 M, NaClO4). The observed rate law has the form: -d[Cr(NH3)5(NCO)2+]/dt = kobs[Cr(NH3)5(NCO)2+] where kobs = a[H+]2{1 + b[H+]2} and ?1 at 55.0°C, a = 0.36 M?1 s?2 and b = 6.9 × 10?3 M?1 s?1. The rate of loss of Cr(NH3)5(NCO)2+ increases with increasing acidity to a limiting value (at [H+] ~ 0.5 M) but the yield of Cr(NH3)63+ decreases with increasing [H+] and increases with increasing temperature. In the kinetic studies the maximum yield of Cr(NH3)63+ was 35% but a synthetic procedure has been developed to give a 60% yield.  相似文献   

15.
A method is introduced by which mass-analysed ion kinetic energy spectra free from Z-discrimination can be obtained for both collisionally activated (CA) and metastable decomposition reactions. The method, performed on a ZAB-E instrument fitted with a collision cell, but applicable also to the ZAB-2F, involves summation of the ‘height resolved’ contributions (formed by beam collimation in the Z-axis and selected by electrostatic deflection of the incident beam) using the signal averaging facility normally available. Representative results (at 8 or 10 keV energy) are given for the CA (Ar target) reactions [CS2]2+ → [CS]+; [CS2]+ → S+ and [CH3OH]+ → [m/z = 12–31]+, and for the metastable reaction [m/z 45]+ → [m/z 29]+ in ethanol.  相似文献   

16.
A highly sensitive and selective method using LC‐ESI‐MS/MS and tandem‐SPE was developed to detect trace amounts of avoparcin (AV) antibiotics in animal tissues and milk. Data acquisition using MS/MS was achieved by applying multiple reaction monitoring of the product ions of [M + 3H]3+ and the major product ions of AV‐α and ‐β at m/z 637 → 86/113/130 and m/z 649 → 86/113/130 in ESI(+) mode. The calculated instrumental LODs were 3 ng/mL. The sample preparation was described that the extraction using 5% TFA and the tandem‐SPE with an ion‐exchange (SAX) and InertSep C18‐A cartridge clean‐up enable us to determine AV in samples. Ion suppression was decreased by concentration rates of each sample solution. These SPE concentration levels could be used to detect quantities of 5 ppb (milk), 10 ppb (beef), and 25 ppb (chicken muscle and liver). The matrix matching calibration graphs obtained for both AV‐α (r >0.996) and ‐β (r >0.998) from animal tissues and milk were linear over the calibration ranges. AV recovery from samples was higher than 73.3% and the RSD was less than 12.0% (n = 5).  相似文献   

17.
The high‐sensitive detection of explosives is of great importance for social security and safety. In this work, the ion source for atmospheric pressure chemical ionization/mass spectrometry using alternating current corona discharge was newly designed for the analysis of explosives. An electromolded fine capillary with 115 µm inner diameter and 12 mm long was used for the inlet of the mass spectrometer. The flow rate of air through this capillary was 41 ml/min. Stable corona discharge could be maintained with the position of the discharge needle tip as close as 1 mm to the inlet capillary without causing the arc discharge. Explosives dissolved in 0.5 µl methanol were injected to the ion source. The limits of detection for five explosives with 50 pg or lower were achieved. In the ion/molecule reactions of trinitrotoluene (TNT), the discharge products of NOx? (x = 2,3), O3 and HNO3 originating from plasma‐excited air were suggested to contribute to the formation of [TNT ? H]? (m/z 226), [TNT ? NO]? (m/z 197) and [TNT ? NO + HNO3]? (m/z 260), respectively. Formation processes of these ions were traced by density functional theory calculations. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

18.
High mass-to-charge ratio ions (> 4000) from electrospray ionization (ESI) have been observed for several proteins, including bovine cytochrome c (M r 12,231) and porcine pepsin (M r 34,584), by using a quadrupole mass spectrometer with an m/z 45,000 range. The ESI mass spectrum for cytochrome c in an aqueous solution gives a charge state distribution that ranges from 12 + to 2 +, with a broad, low-intensity peak in the mass-to-charge ratio region corresponding to the [M + H]+ ion. the negative ion ESI mass spectrum for pepsin in 1% acetic acid solution shows a charge state distribution ranging from 7? to 2?. To observe the [M - H]? ion, harsher desolvation and interface conditions were required. Also observed was the abundant aggregation of the protens with average charge states substantially lower than observed for their monomeric counterparts. The negative ion ESI mass spectrum for cytochrome c in 1–100 mM NH4OAc solutions showed greater relative abundances for the higher mass-to-charge ratio ions than in acuidic solutions, with an [M - H]? ion relative abundance approximately 50% that of the most abundant charge state peak. The observation that protein aggregates are formed with charge states comparable to monomeric species (at fower mass-to-charge ratios) suggests that the high mass-to-charge ratio monomers may be formed by the dissociation of aggregate species. The observation of low charge state and aggregate molecular ions concurrently with highly charged species may serve to support a variation of the charged residue model, originally described by Dole and co-workers (Dole, M., et al. J. Chem. Phys. 1968, 49, 2240; Mack, L. L., et al. J. Chem. Phys. 1970, 52, 4977) which involves the Coulombically driven formation of either very highly solvated molecular ions or lower ananometer-diameter droplets.  相似文献   

19.
o-Phthalic acid is proposed as a standard substance for buffer solutions of known hydrogen ion concentration (I ? 0.2 M KCl, p[H+] = 3.0–5.4, 25°C). Its crystallinity, purity and slightly wide buffer range afford advantages over acetic acid. Empirical relationships between measured pH (pHm) and calculated [H+] were derived for sequences of buffer solutions at several ionic strengths: pHm - Mp[H+] + C. These calibration lines were parallel and of unit slope as required by theory. A table of p[H+] values for o-phthalic acid buffer solutions at I = 0.1 M (KCl) is presented and the method of calculation of p[H+] values for a buffer series generated by additions of potassium hydroxide is outlined.  相似文献   

20.
Cluster ions such as [Cat+X+nM](+) (n = 0-4); [Cat-H+nM](+) (n = 1-3); and [2(Cat-H)+X+nM](+) (n = 0-2), where Cat, X, and M are the dication, anion, and neutral salt (CatX(2)), respectively, are observed in electrospray ionization (ESI) mass spectrometry of relatively concentrated solutions of diquat and paraquat. Collision-induced dissociation (CID) reactions of the clusters were observed by tandem mass spectrometry (MS/MS), including deprotonation to form [Cat-H](+), one-electron reduction of the dication to form Cat(+.), demethylation of the paraquat cation to form [Cat-CH(3)](+), and loss of neutral salt to produce smaller clusters. The difference in acidity and reduction power between diquat and paraquat, evaluated by thermodynamical estimates, can rationalize the different fractional yields of even-electron ([Cat-H](+) and its clusters) and odd-electron (mostly Cat(+)) ions in ESI mass spectra of these pesticides. The [Cat+n. Solv](2+) doubly charged cluster ions, where n 相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号